Tryptophan-136 in Subunit II of Cytochrome bo3 ... - ACS Publications

Helsinki Bioenergetics Group, Department of Medical Chemistry, Institute of Biomedical Sciences, and Biocentrum Helsinki,. P.B. 8, UniVersity of Helsi...
0 downloads 0 Views 131KB Size
11806

Biochemistry 1998, 37, 11806-11811

Tryptophan-136 in Subunit II of Cytochrome bo3 from Escherichia coli May Participate in the Binding of Ubiquinol† Jixiang Ma,‡,§ Anne Puustinen,| Ma˚rten Wikstro¨m,| and Robert B. Gennis*,‡ Department of Biochemistry, UniVersity of Illinois, 600 South Mathews Street, Urbana, Illinois 61801, and Helsinki Bioenergetics Group, Department of Medical Chemistry, Institute of Biomedical Sciences, and Biocentrum Helsinki, P.B. 8, UniVersity of Helsinki, Helsinki, FI-00014 Finland ReceiVed May 1, 1998; ReVised Manuscript ReceiVed June 23, 1998

ABSTRACT: In the cytochrome c oxidases, the role of subunit II is to provide the electron entry site into the enzyme. This subunit contains both the binding site for the substrate, cytochrome c, and the CuA redox center, which is initially reduced by cytochrome c. Cytochrome bo3 and other quinol oxidases that are members of the heme-copper oxidase superfamily have a homologous subunit II, but the CuA site is absent, as is the docking site for cytochrome c. Speculation that subunit II in the quinol oxidases may also be important as an electron entry site is supported by the demonstration several years ago that a photoreactive substrate analogue, azido-Q, covalently labeled subunit II in cytochrome bo3. In the current work, a sequence alignment of subunit II of heme-copper quinol oxidases is used as a guide to select conserved residues that might be important for the binding of ubiquinol to cytochrome bo3. Results are presented for point mutants in 24 different residue positions in subunit II. The membrane-bound enzymes were examined by optical spectroscopy and by determining the activity of ubiquinol-1 oxidase. In each case, the Km for ubiquinol-1 was determined as a measure of possible perturbation to a quinol binding site. The only mutant that had a noticeably altered Km for ubiquinol-1 was W136A, in which the Km was about sixfold increased. Thus, W136 may be at or close to a substrate (ubiquinol)-binding site in cytochrome bo3. In the cytochrome c oxidases, the equivalent tryptophan (W121 in Paracoccus denitrificans) has been identified as the “electron entry site”.

Cytochrome bo3 from Escherichia coli is a member of the heme-copper oxidase superfamily that includes the cytochrome c oxidases (1). These enzymes all catalyze the 4-electron reduction of dioxygen to water and are proton pumps, generating a protonmotive force (2). Cytochrome bo3, however, does not utilize cytochrome c as a substrate but instead catalyzes the 2-electron oxidation of ubiquinol-8 in the cytoplasmic membrane of E. coli. The enzyme contains one copy each of four subunits (3). Subunits I, II, and III are homologues to those found in other heme-copper oxidases, whereas subunit IV is unrelated to the subunits of the cytochrome c oxidases (4). Because the structure of the cytochrome c oxidases from bovine heart mitochondria (5, 6) and from the bacterium Paracoccus denitrificans (7, 8) have been determined by X-ray crystallography, a reasonable model of the structure of subunits I, II, and III of cytochrome bo3 can be made (9). Subunit I, which has the highest sequence homologies among the heme-copper oxidases, contains both the active site (a heme-copper bimetallic center) and a low-spin heme component whose function is to transfer electrons to the heme-copper center (1, 2). The † This work was supported by a grant from the U.S. Department of Energy, DEFG02-87ER13716 (R.B.G.), and by the Academy of Finland and the Sigrid Juse´lius Foundation (M.W.). * Corresponding author: Telephone: 217-333-9075. Fax: 217-2448641. E-mail: [email protected]. ‡ University of Illinois. § Current address: NIEHS, 111 Alexander Drive, MD D2-01, Research Triangle Park, NC 27709, USA. | University of Helsinki.

proton pump machinery is also contained within subunit I. In the cytochrome c oxidases, one role of subunit II is clearly to provide electrons to subunit I. The binding site for cytochrome c is located on the surface of a large hydrophilic COOH-terminal domain of subunit II (10, 11), which is anchored to the membrane by two NH2-terminal transmembrane helices. Also contained within subunit II is the CuA redox center, which accepts electrons from cytochrome c and passes them along to the low-spin heme within subunit I (2, 6, 7). Subunit II of cytochrome bo3 shares the same general structure as that of the cytchrome c oxidases (12-14) but lacks both the binding site of cytochrome c and the CuA center. The structure of the COOH-terminal hydrophilic domain has been determined by X-ray crystallography (14), and is clearly closely related to the homologous domains of the cytochrome c oxidases (6, 7). Indeed, by site-directed mutagenesis, the CuA site has been restored to the subunit II domain and to cytochrome bo3 (15). We therefore postulate that the role of subunit II in the quinol oxidases may be analogous to that of the equivalent subunit in the cytochrome c oxidases, i.e., providing electrons to the metal centers within subunit I. However, whereas cytochrome c is watersoluble and binds to subunit II from the aqueous phase, ubiquinol-8 is hydrophobic and is confined to the membrane. Clearly, if subunit II of cytochrome bo3 does contain a ubiquinol-binding site, it cannot be a simple modification of the analogous binding site for cytochrome c.

S0006-2960(98)00997-0 CCC: $15.00 © 1998 American Chemical Society Published on Web 08/01/1998

Ubiquinol-Binding Site of Cytochrome bo3 Subunit II has two transmembrane helices, both in contact with subunit I, and the hydrophilic domain sits on the top of subunit I (on the periplasmic surface for cytochrome bo3) and forms a cap over the heme centers that is essentially at the membrane interface (6, 7). Support for a role for subunit II in the binding of ubiquinol in cytochrome bo3 is provided by the use of a photoreactive substrate analogue, azido-Q (16, 17). Upon photolysis in the presence of purified cytochrome bo3, azido-Q is covalently linked to subunit II (16). To further localize the binding site for ubiquinol within subunit II, we tried two approaches. In one, we used the covalent attachment of azido-Q, presumably at a ubiquinolbinding site, to identify a peptide likely to participate in ubiquinol binding. By mass spectroscopy, a trypsin fragment consisting of residues 165-178 of Subunit II was determined to be linked to azido-Q (17). In the second, parallel approach, we used site-directed mutagenesis, directed at selected residues within subunit II, to determine which residues are functionally important and might be involved in ubiquinol binding. Here, we report the results of the mutagenesis study. Of the 24 different sites within subunit II that we altered by mutagenesis, only 1 (residue W136) appeared to be functionally important and was thus potentially a candidate for direct involvement in the binding of ubiquinol. MATERIALS AND METHODS Materials. The restriction enzymes KpnI, SalI, HindIII, and MspI were purchased from Bethesda Research Laboratories (BRL) . The restriction enzyme HhaI, T4 DNA ligase, and T4 DNA kinase were from New England Biolabs. T4 DNA polymerase, exonuclease III, and sequenase kits were obtained from U.S. Biochemicals. The nucleotides dATP, dGTP, dCTP, dTTP, and 5′Me-dCTP were from Pharmacia LKB Biotechnology. ATP and phenylmethylsulfonyl fluoride (PMSF) were from Sigma. Oligonucleotides used for mutation or sequencing were obtained from the Biotechnology Center at the University of Illinois. Ubiquinone-1 was a gift from Hoffmann-LaRoche AG (Basel, Switzerland). Mutagenesis. Single-stranded M13 DNA was prepared according to the procedures of Messing and colleagues (18). Mutagenesis was performed as described (19), with the following modifications. The mutagenic primer (∼75 pmol) was 5′-phosphorylated with 1 unit of T4 nucleotide kinase in a 30-µL reaction containing 100 mM Tris-HCl, 10 mM MgCl2, 7 mM dithiothreitol, and 1 mM ATP, at pH 8.0. After incubation at 37 °C for 30 min, the reaction was stopped by heating to 70 °C for 10 min. We then mixed 2 µL of 0.1 µg/µL M13 DNA template with 2 µL of the above reaction mixture, added 5 µL of H2O and 1 µL of 10× React 2 (BRL) buffer (500 mM Tris-HCl, 100 mM MgCl2, and 500 mM NaCl, pH 8.0) and incubated the mixture at 65 °C for 5 min. After the mixture cooled at room temperature for 15 min, the second DNA strand was synthesized by adding to the same tube 8 µL of 10× React 1 (BRL) buffer (500 mM TrisHCl and 100 mM MgCl2, pH 8.0), 19 µL of mixed dNTPs solution (1.33 mM each of dATP, dGTP, dTTP, and 5-methylated-dCTP, plus 2.67 mM ATP), 1 µL of T4 DNA ligase, and 5 units of T4 DNA polymerase. Water was added to the mixture to bring the total volume to 100 µL. After incubation for 90 min at 37 °C, the extension and ligation reactions were terminated by incubation at 70 °C for 10 min.

Biochemistry, Vol. 37, No. 34, 1998 11807 The parent DNA template was nicked and digested with 1 µL each of MspI and HhaI for 45 min, followed by 1 µL of exonuclease III at 37 °C for 45 min. The mixture was heated to 70 °C for 10 min, and then transformed into strain Jmr-, which allows methylated DNA to be propagated. Once the desired mutations were identified on M13 phage by single-stranded DNA sequencing, the fragments containing the mutations were cloned back into cyo operon in plasmid pMC31 (20). After subcloning, all mutations were confirmed by double-stranded DNA sequencing. GL101 (20) was used to propagate both wild-type and mutant plasmids. Complementation Analysis of the Mutant Plasmids. Genetic complementation was conducted with the procedure previously described (20). The host strain used for complementation analysis was either RG129 (21) or GO105 (22), which cannot grow aerobically on nonfermentable substrates because of mutations in both the cyo and cyd operons of the chromosomes. Overnight aerobic growth on succinatelactate-minimal plates at 37 °C was the criterion used to determine the presence of a functional cytochrome bo3. Cell Growth and Membrane Preparation. Strain GO105 (22) was used as the host strain for further characterization. Typically, a single colony harboring a mutant plasmid or the wild type was inoculated into 10 mL of liquid LB supplemented with ampicillin (100 µg/mL) and kanamycin (50 µg/mL). After overnight aerobic growth, the 10-mL culture was used to inoculate 1 L of LB with 0.3% D,L-lactate, 200 µM CuSO4, and the antibiotics mentioned above. The cells were grown at 37 °C in 2-L baffled flasks with vigorous shaking. After the cell density was increased by ∼70 Klett units (determined with a red filter), the cells were harvested by centrifugation and then washed with 10 mM Tris-HCl, pH 8.0. The washed cells from a 1-L culture were resuspended in 20 mL of buffer containing 10 mM TrisHCl, 0.5 mM PMSF, leupeptin 0.5 µg/mL, 5 mM MgSO4, and DNAse 4 µg/mL. The cell suspension was passed twice through a French pressure cell at 1200 psi (8.28 MPa), and cell debris and most of the outer membrane were removed by centrifugation at 9000g for 30 min, followed by a second centrifugation at 25 000g for 30 min. The supernatant was centrifuged at 160 000g in a 60-Ti rotor for 1.5 h. The red cytoplasmic membranes were collected and homologized in 2 mL of ice-cold 50 mM potassium phosphate buffer (pH 7.5) with leupeptin (0.5 µg/mL) for subsequent characterization. Oxidase ActiVity Assay. Ubiquinol-1 oxidase activity at 37 °C was measured with a model 53 YSI oxygen electrode (Yellow Springs Instrument Co., Yellow Springs, OH) and a temperature-controlled 1.8-mL electrode chamber (Gibson). The buffer for this assay on membrane samples was 50 mM Tris and 1 mM EDTA, pH 7.4, containing 2 mM dithioerythritol and 150 µM ubiquinone-1. For assays with the purified oxidase, the buffer contained 50 mM Hepes, 1 mM EDTA, 0.01% dodecyl maltoside, 2 mM dithiothreitol, and 150 µM ubiquinol-1 (pH 7.4). The concentration of O2 in the air-saturated buffers at this temperature was assumed to be 250 µM. The reaction was initiated by injecting 5 µL of appropriately diluted membrane preparations or pure enzymes. The background drift prior to the addition of oxidases was subtracted from the results.

11808 Biochemistry, Vol. 37, No. 34, 1998

Ma et al.

FIGURE 1: Sequence alignment of subunit II of the bacterial quinol oxidases. Residues that are totally conserved are boxed. The sequences are from the quinol oxidases from E. coli (4), Acetobacter aceti (35), Bacillus subtilis (36), and P. denitrificans (37).

Protein Purification: Wild type and W136A mutants were purified according to the published procedure of Minghetti et al. (23). Proton Pumping Assay. Proton pumping was measured as previously described (24, 25). Miscellaneous. Protein concentration was determined with the bicinchoninic acid method (Pierce Chemical Co.). Optical spectroscopy was carried with a DW2000 UV-Vis spectrometer (SLM Instruments). The heme composition of purified mutant cytochrome bo3 oxidases was analyzed as previously described (26). RESULTS Site-directed mutagenesis was used to screen a large set of residues within subunit II of cytochrome bo3 for structural or functional importance. Figure 1 shows a sequence alignment of four ubiquinol oxidases within the hemecopper superfamily. Excluded from this list are the quinol oxidases from the archaebacteria, which have unique properties and are grouped in a separate phylogenetic category (27-

29). As Figure 1 shows, a number of residues are conserved within the quinol oxidases; some of these are also highly conserved in the cytochrome c oxidases (not shown). Although residues common to the quinol oxidases are scattered throughout the sequences, three clusters are apparent, roughly E89 to I99, D135 to Q145, and Y184 to M189 (E. coli numbering). By comparison with the structures of the cytochrome c oxidases (5, 7, 8), we determined the first cluster (E89 to I99) to be located near the cytoplasmic side of the membrane within the second transmembrane helix. Therefore, those residues are not likely to be involved in the binding of ubiquinol, which is expected to be oxidized at the periplasmic surface of the membrane. Protons from the oxidation of ubiquinol are released into the bacterial periplasm (9, 30), and the metal centers are located toward this side of the membrane (6, 7). Therefore, we selected the targets for mutagenesis from residues within the periplasmic COOH-terminal domain, i.e., encompassing the other two clusters of conserved residues: D135 to Q145 and Y184 to M189 (Figure 1). The selected residues do not

Ubiquinol-Binding Site of Cytochrome bo3

Biochemistry, Vol. 37, No. 34, 1998 11809

Table 1: Summary of Properties of Site-Directed Mutants of Subunit II of Cytochrome bo3

mutant

complementation

relative oxidase specific activitya, %

relative expressionb, %

KM, µM

WT D135N W136A K137M W138V F139A F140A Y142F Y142V P143A E144Q Q145A T149A P156A V170A P177A Y184F Y184V A185V M186A G188A M189A Y203A Y203F G205A Y210F G212A M218A

yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes

100 118 72 95 63 72 95 95 63 105 120 97 95 97 109 118 94 106 102 104 103 117 98 109 110 100 76 88

100 90 99 93 52 88 106 83 56 85 64 84 104 83 64 76 67 69 103 108 86 93 101 115 79 115 81 90

38 56 242 44 45 39 59 48 31 86 67 73 58 37 66 81 34 43 59 75 52 54 63 31 65 61 45 59

a The specific activity of the wild-type oxidase from the parent plasmid was taken as 100%. Ubiquinol-1 oxidase activity was expressed per mole of CO-binding heme o3. All assays were performed with membrane preparations. b Determined by the relative CO-binding units in membrane samples, using an absorptivity of 135 mM-1cm-1.

include all of those that are conserved in the quinol oxidases in this region but are spread through the region from ∼130 to 230. The residues were also selected without regard for their positions in the structure of the hydrophilic domain of subunit II of cytochrome bo3 (14) to provide a potentially wider range of phenotypes. In addition, several of the residues selected are not fully conserved. Table 1 summarizes the properties of the mutants. All of the mutants are expressed to approximately the same amount and are functional, both by genetic complementation and by biochemical assay. Since a subtle perturbation to the ubiquinol binding site might not be apparent in a single measurement of catalytic velocity, we determined the dependence of activity on the concentration of ubiquinol-1. The data in all cases could be interpreted in terms of a Km for ubiquinol-1. As shown in Table 1, only the W136A mutation had a Km significantly different from that of the wild-type control. The W136A mutation had a sixfold increase of the Km of the oxidase for ubiquinol-1, considerably more than the fluctuations in Km values observed for the other mutant oxidases. The W136A oxidase was solubilized and purified for further analysis. Spectroscopic studies, including the reducedminus-oxidized spectrum and the CO-reduced-minus-reduced spectrum, showed no difference from the results for the wildtype oxidase. There is thus no indication of any perturbation to either the low-spin heme or the heme-copper binuclear center. The dependence of the ubiquinol-1 oxidase activity

FIGURE 2: A Lineweaver-Burk plot showing the dependence of the steady-state kinetics of the ubiquinol-1 oxidase activity on the concentration of ubiquinol-1 for the purified wild type (WT) and W136A mutant of cytochrome bo3. The data were analyzed to obtain the Km values, which are threefold higher for the W136A mutant. The Km values in Table 1 were obtained by using membrane-bound enzymes, in which case there is a sixfold increase due to the W136A mutation. Table 2: Proton Pumping Stoichiometry of Selected Mutants of Cytochrome bo3 H+/emutant

pH 6.0

pH 7.0

pH 8.0

WT W136A W138V Y142F

1.6-2 1.7-1.9 1.7-1.85 1.5-2

1.3-1.6 1.5-1.8 1.3-1.7 1.4-1.5

1-1.3 1 0.9 1-1.3

was also determined by using the purified wild-type and W136A mutant oxidases. Each has an identical maximal turnover number (Vmax), ∼1200 electrons per second, but the Km for ubiquinol-1 is threefold higher for the W136A mutant (Figure 2). The reason the influence of the mutation is less after purification (threefold increase in Km) than in the membrane-bound form (sixfold increase) is not known, but both increases in Km were consistent, as determined with several independent preparations of the enzyme. Although no other mutants showed a significant change in the Km for quinol, two mutations displayed a lower specific activity than that of the wild type: W138V and Y142V (but not Y142F) (Table 1). In each case, the strains harboring these mutant oxidases grew slowly, and the specific quinol oxidase activities were ∼60% of that of the wild type. However, the Km for ubiquinol-1 was not altered in either case. As with W136A, neither W138V nor Y142F showed any spectroscopic perturbations. Finally, proton pumping assays (at pH 6.0, 7.0, and 8.0) were performed with cells containing the W136A mutant and with cells containing W138V or Y142F (Table 2). No difference from wild-type behavior was observed. These mutations have no influence on the proton pumping stoichiometry of cytochrome bo3 or on its pH-dependence (25). DISCUSSION Our motivation in this study was to obtain information about the location of the ubiquinol-binding site(s) of cytochrome bo3, given the finding that a photoreactive ubiquinol analogue, azido-Q, covalently labels subunit II (16). Presumably, some of the residues making up the ubiquinol-

11810 Biochemistry, Vol. 37, No. 34, 1998

FIGURE 3: Structure of the hydrophilic COOH-terminal domain of subunit II of cytochrome bo3 from E. coli (14). The location of tryptophan-136 is indicated along with the peptide that has been implicated in the binding site for ubiquinol (17). Also shown is the reconstructed site where the bimetallic CuA site is located. This redox site is not present in wild-type cytochrome bo3 but is present in the related cytochrome c oxidases.

binding site would be common to the set of related quinol oxidases. The sequence alignment of subunit II of the ubiquinol oxidases indicates a number of conserved residues, some of which are also common to the much larger set of the related cytochrome c oxidases. Residue C25 is one of the residues conserved in the set of four quinol oxidases. We previously showed (31) that in the E. coli oxidase the posttranslational processing results in cleavage of the protein C25 and covalent addition of lipids to this residue, anchoring the modified NH2-terminus of the subunit to the membrane. However, this NH2-terminal modification is not important for the interaction of the enzyme with ubiquinol (31). Of the 27 mutants examined in the current study, only the W136A mutation resulted in a significant, albeit modest, effect on the Km of ubiquinol-1. We conclude that W136 is probably located at or very close to a ubiquinol-binding site in cytochrome bo3. Concurrent with this work, we also determined that the azido-Q photolabel covalently binds to peptide residues 165-178 in subunit II, implicating some portion of this peptide in the binding of ubiquinol (17). Figure 3 shows a model of the structure of cytochrome bo3, highlighting both this peptide as well as W136. Clearly, these are close to each other, which suggests a ubiquinolbinding site near the interface between subunits I and II. Since quinol is oxidized by the metal centers in subunit I, and since the binding site must be accessed from the lipid bilayer, this location for the binding site is plausible. Some evidence suggests that cytochrome bo3 contains two sites for binding ubiquinol (32, 33): One site has high affinity and appears not to readily exchange with free ubuqinol (33, 34), whereas the second site has lower affinity and appears to be a traditional substrate-binding site. Presumbly, W136A perturbs the low-affinity site, which would be expected to result in an altered Km for the substrate. A model of this site has been generated by molecular modeling, docking a ubiquinone close to both W136 and the peptide 165-178 (17). Four of the seven residues that were predicted to be within 5 Å, but further than 3 Å, from the docked quinone were coincidentally altered by mutagenesis in the current work: D135, V170, Y210, and M218. The mutations at these sites (Table 1) did not alter the Km for ubiquinol-1,

Ma et al. however, which suggests they are not directly involved in or required for the binding of ubiquinol, presumably because they are too far away. Additional sites will be targeted for mutagenesis to further examine the model. The current data are also consistent with recent studies from the group of Anraku (Sato-Watanabe et al., unpublished) in which mutants resistant to quinol analogues were isolated and mapped. Given that these mutations were all within the hydrophilic domain of subunit II, the work independently suggests a quinol-binding site in the region near W136 and the CuA domain. Of particular interest, the residue equivalent to W136 in the cytochrome c oxidase from P. denitrificans, W121, has been shown to be functionally important in mediating electron transfer from cytochrome c to CuA (11). W121 is located within the binding domain for cytochrome c on the surface of subunit II. A significant role for W136 in the substrate-binding site of cytochrome bo3 would represent an interesting functional evolution from the cytochrome c oxidases to the quinol oxidases. ACKNOWLEDGMENT We thank Betsy Furlong Nickels for assistance with preparing Figures 1 and 3 and Tatsushi Mogi for providing data prior to publication. REFERENCES 1. Garcia-Horsman, J. A., Barquera, B., Rumbley, J., Ma, J., & Gennis, R. B. (1994) J. Bacteriol. 176, 5587-5600. 2. Ferguson-Miller, S., & Babcock, G. T. (1996) Chem. ReV. 7, 2889-2907. 3. Rumbley, J. N., Nickels, E. F., & Gennis, R. B. (1997) Biochim. Biophys. Acta 1340, 131-142. 4. Chepuri, V., Lemieux, L. J., Au, D. C.-T., & Gennis, R. B. (1990) J. Biol. Chem. 265, 11185-11192. 5. Tsukihara, T., Aoyama, H., Yamashita, E., Takashi, T., Yamaguichi, H., Shinzawa-Itoh, K., Nakashima, R., Yaono, R., & Yoshikawa, S. (1996) Science 272, 1136-1144. 6. Tsukihara, T., Aoyama, H., Yamashita, E., Tomizaki, T., Yamaguchi, H., Shinzawa-Itoh, K., Nakashima, T., Yaono, R., & Yoshikawa, S. (1995) Science 269, 1069-1074. 7. Iwata, S., Ostermeier, C., Ludwig, B., & Michel, H. (1995) Nature 376, 660-669. 8. Ostermeier, C., Harrenga, A., Ermler, U., & Michel, H. (1997) Proc. Natl. Acad. Sci. U.S.A. 94, 10547-10553. 9. Trumpower, B. L., & Gennis, R. B. (1994) Annu. ReV. Biochem. 63, 675-716. 10. Witt, H., Malatesta, F., Nicoletti, F., Brunori, M., & Ludwig, B. (1998) FEBS Lett. 251, 367-373. 11. Witt, H., Malatesta, P., Nicoletti, F., Brunori, M., & Ludwig, B. (1998) J. Biol. Chem. 273, 5132-5136. 12. Calhoun, M. W., Thomas, J. W., & Gennis, R. B. (1994) Trends Biochem. Sci. 19, 325-330. 13. Gohlke, U., Warne, A., & Saraste, M. (1997) EMBO J. 16, 1181-1188. 14. Wilmanns, M., Lappalainen, P., Kelly, M., Sauer-Eriksson, E., & Saraste, M. (1995) Biochemistry 92, 11955-11959. 15. van der Oost, J., Pappalainen, P., Musacchio, A., Warne, A., Lemieux, L., Rumbley, J., Gennis, R. B., Aasa, R., Pascher, T., Malmstro¨m, B. G., & Saraste, M. (1992) EMBO J. 11, 3209-3217. 16. Welter, R., Gu, L.-Q., Yu, L., Yu, C.-A., Rumbley, J., & Gennis, R. B. (1994) J. Biol. Chem. 269, 28834-28838. 17. Tsatsos, P. H., Reynolds, K., Nickels, E. F., He, D.-Y., Yu, C.-A., & Gennis, R. B. (1998) Biochemistry, 37, 9884-9888. 18. Yanisch-Perron, C., Vieira, J., & Messing, J. (1985) Gene 33, 103-119.

Ubiquinol-Binding Site of Cytochrome bo3 19. Vandeyar, M. A., Weiner, M. P., Hutton, C. J., & Batt, C. A. (1988) Gene 65, 129-133. 20. Lemieux, L. J., Calhoun, M. W., Thomas, J. W., Ingledew, W. J., & Gennis, R. B. (1992) J. Biol. Chem. 267, 21052113. 21. Au, D. C.-T., & Gennis, R. B. (1987) J. Bacteriol. 169, 32373242. 22. Calhoun, M. W., Lemieux, L. J., Thomas, J. W., Hill, J. J., Goswitz, V. C., Alben, J. O., & Gennis, R. B. (1993) Biochemistry 32, 13254-13261. 23. Minghetti, K. C., Goswitz, V. C., Gabriel, N. E., Hill, J. J., Barassi, C., Georgiou, C. D., Chan, S. I., & Gennis, R. B. (1992) Biochemistry 31, 6917-6924. 24. Thomas, J. W., Puustinen, A., Alben, J. O., Gennis, R. B., & Wikstro¨m, M. (1993) Biochemistry 32, 10923-10928. 25. Verkhovskaya, M., Verkhovsky, M., & Wikstro¨m, M. (1992) J. Biol. Chem. 267, 14559-14562. 26. Puustinen, A., & Wikstro¨m, M. (1991) Proc. Natl. Acad. Sci. U.S.A. 88, 6122-6126. 27. Castresana, J., Lu¨bben, M., & Saraste, M. (1995) J. Mol. Biol. 250, 202-210. 28. Castresana, J., & Saraste, M. (1995) Trends Biochem. Sci. 20, 443-448.

Biochemistry, Vol. 37, No. 34, 1998 11811 29. Lu¨bben, M. (1995) Biochim. Biophys. Acta 1229, 1-22. 30. Puustinen, A., Finel, M., Virkki, M., & Wikstro¨m, M. (1989) FEBS Lett. 249, 163-167. 31. Ma, J., Katsonouri, A., & Gennis, R. B. (1997) Biochemistry 36, 11298-11303. 32. Musser, S. M., Stowell, M. H. B., Lee, H. K., Rumbley, J. N., & Chan, S. I. (1997) Biochemistry 36, 894-902. 33. Sato-Watanabe, M., Mogi, T., Ogura, T., Kitagawa, T., Miyoshi, H., Iwamura, H., & Anraku, Y. (1994) J. Biol. Chem. 269, 28908-28912. 34. Puustinen, A., Verkhovsky, M. I., Morgan, J. E., Belevich, N. P., & Wikstro¨m, M. (1996) Proc. Natl. Acad. Sci. U.S.A. 93, 1545-1548. 35. Fukaya, M., Tayama, K., Tamaki, T., Ebisuya, H., Okumura, H., Kawamura, Y., Horinouchi, S., & Beppu, T. (1993) J. Bacteriol. 175, 4307-4314. 36. Richter, O.-M. H., Tao, J.-S., Turba, A., & Budwig, B. (1994) J. Biol. Chem. 269, 23079-23086. 37. Santana, M., Kunst, F., Hullo, M. F., Rapoport, G., Danchin, A., & Glaser, P. (1992) J. Biol. Chem. 267, 10225-10231. BI9809977